Explicative note to this paper - McGill Physicsgang/eprints/eprintLovejoy/neweprint/climate... ·...

36
Page 1 Explicative note to this paper 1 This paper was submitted to BAMS on June 27, 2012, it was accepted for 2 publication by editor C. Deser on Dec. 27, 2012 and retracted by the senior BAMS editor 3 J. Rosenfeld on Oct. 24, 2013. The reason given was that a much shorter review of 4 similar material (29% of the length) had been published in the weekly magazine EOS 5 (Lovejoy, S., 2013: What is Climate?: EOS, 94, No. 1, 1 January 2013, p1-2.). Lovejoy 6 was therefore accused of attempting “concurrent dual publication”. 7 The background was that at the moment of submission, the draft paper was posted 8 on Lovejoy’s personal web site. It rapidly gained notoriety in the blogosphere and came 9 to the attention of the EOS editors R. Pielke Sr. and B. Richman. In August 2012, this 10 much shorter review was explicitly commissioned by EOS to cover similar material but 11 in much shorter form and for a quite a different audience. The practice of EOS 12 commissioning short reviews of new work covered at greater length in other journals is 13 common and there was certainly no attempt at dual publication. More details on this 14 unfortunate episode may be found at in the chronology document 15 (http://www.physics.mcgill.ca/~gang/eprints/eprintLovejoy/neweprint/ChronologyBAMS 16 .EOS.25.10.13.pdf). 17 18

Transcript of Explicative note to this paper - McGill Physicsgang/eprints/eprintLovejoy/neweprint/climate... ·...

  • Page 1

    Explicative note to this paper 1  

    This paper was submitted to BAMS on June 27, 2012, it was accepted for 2  publication by editor C. Deser on Dec. 27, 2012 and retracted by the senior BAMS editor 3  J. Rosenfeld on Oct. 24, 2013. The reason given was that a much shorter review of 4  similar material (29% of the length) had been published in the weekly magazine EOS 5  (Lovejoy, S., 2013: What is Climate?: EOS, 94, No. 1, 1 January 2013, p1-2.). Lovejoy 6  was therefore accused of attempting “concurrent dual publication”. 7  

    The background was that at the moment of submission, the draft paper was posted 8  on Lovejoy’s personal web site. It rapidly gained notoriety in the blogosphere and came 9  to the attention of the EOS editors R. Pielke Sr. and B. Richman. In August 2012, this 10  much shorter review was explicitly commissioned by EOS to cover similar material but 11  in much shorter form and for a quite a different audience. The practice of EOS 12  commissioning short reviews of new work covered at greater length in other journals is 13  common and there was certainly no attempt at dual publication. More details on this 14  unfortunate episode may be found at in the chronology document 15  (http://www.physics.mcgill.ca/~gang/eprints/eprintLovejoy/neweprint/ChronologyBAMS16  .EOS.25.10.13.pdf). 17   18  

  • Page 2

    19  

    The  climate  is  not  what  you  expect  20  

    S.  Lovejoy1,  D.  Schertzer2    21  

    1Physics,  McGill  University  22  

    3600  University  st.,  23  

    Montreal,  Que.,  Canada  24  

    Email:  [email protected]  25  

     26  

    2CEREVE  27  

    Université  Paris  Est,  28  

    6-‐8,  avenue  Blaise  Pascal  29  

    Cité  Descartes  30  

    77455  MARNE-‐LA-‐VALLEE  Cedex,  France  31  

       32  

  • Page 3

    The  climate  is  not  what  you  expect  33  

    Capsule  Summary:  34  

    Contrary  to  popular  ideas  about  the  climate,  what  you  really  expect  is  35  

    macroweather.    On  scales  of  a  human  generation,  the  climate  is  what  you  get.  36  

    Abstract:  37  

    Prevailing   definitions   of   climate   are   not  much   different   from   “the   climate   is  38  

    what  you  expect,  the  weather  is  what  you  get”.    Using  a  variety  of  sources  including  39  

    reanalyses   and   paleo   data,   and   aided   by   notions   and   analysis   techniques   from  40  

    Nonlinear   Geophysics,   we   argue   that   this   dictum   is   fundamentally   wrong.     In  41  

    addition   to   the  weather   and   climate,   there   is   a   qualitatively   distinct   intermediate  42  

    regime  extending  over  a  factor  of  ≈  1000  in  scale.    For  example,  mean  temperature  43  

    fluctuations   increase   up   to   about   5   K   at   10   days   (the   lifetime   of   planetary  44  

    structures),   then   decrease   to   about   0.2   K   at   30   years,   and   then   increase   again   to  45  

    about  5  K  at  glacial-‐interglacial  scales.  46  

    Both  deterministic  GCM’s  with  fixed  forcings  (“control  runs”)  and  stochastic  47  

    turbulence-‐based  models   reproduce   the   first   two   regimes,   but   not   the   third.     The  48  

    middle   regime   is   thus   a   kind   of   “macroweather”   not   “high   frequency   climate”.  49  

    Averaging   macroweather   over   periods   increasing   to   ≈   30   yrs   yields   apparently  50  

    converging   values:   macroweather   is   “what   you   expect”.     Macroweather   averages  51  

    over  ≈  30  years  have  the  lowest  variability,  they  yield  well  defined  “climate  states”  52  

    and   justify   the   otherwise   ad   hoc   “climate   normal”   period.     However,   moving   to  53  

  • Page 4

    longer   periods,   these   states   increasingly   fluctuate:   just   as   with   the   weather,   the  54  

    climate   changes   in   an   apparently   unstable   manner;   the   climate   is   not   what   you  55  

    expect.     Similarly,   we  may   categorize   climate   forcings   according   to  whether   their  56  

    fluctuations  decrease  or  increase  with  scale  and  this  has  important  implications  for  57  

    GCM’s  and  for  climate  change  and  climate  predictions.  58  

    1. Introduction 59  

    In his monumental “Climate: Past, Present, and Future” Horace Lamb argued that 60  

    the early scientific view was “climate as constant” [Lamb, 1972]. Reflecting this, in 61  

    1935 the International Meteorological Organization adopted 1901-1930 as the “climatic 62  

    normal period”. Following the post war cooling, this view evolved: for example the 63  

    official American Meteorological Society glossary [Huschke, 1959] defined the climate 64  

    as “the synthesis of the weather” and then “…the climate of a specified area is 65  

    represented by the statistical collective of its weather conditions during a specified 66  

    interval of time (usually several decades)”. Although this new definition in principle 67  

    allows for climate change, the period 1931-1960 soon became the new “normal”, the ad 68  

    hoc 30 year duration became entrenched, today 1961-1990 is commonly used. Mindful 69  

    of the extremes, Lamb warned against reducing the climate to just “average weather”, 70  

    while viewing the climate as “…the sum total of the weather experienced at a place in the 71  

    course of the year and over the years”, [Lamb, 1972]. 72  

    Lamb’s  essentially  modern  view  allows   for   the  possibility  of  climate  change  73  

    and  is  closely  captured  by  the  popular  expression:  “The  climate  is  what  you  expect,  74  

    the  weather   is  what  you  get”  (the   character  Lazurus  Long   in   [Heinlein, 1973],   but  75  

    often   attributed   to   Mark   Twain).     It   is   also   close   to   the   US   National   Academy   of  76  

  • Page 5

    Science  definition:  “Climate   is   conventionally  defined  as   the   long-‐term  statistics  of  77  

    the  weather…”   [Committee on Radiative Forcing Effects on Climate, 2005] which  78  

    improves   on   the   “the   climate   is  what   you   expect”   idea   only   a   little   by   proposing:  79  

    “…to   expand   the   definition   of   climate   to   encompass   the   oceanic   and   terrestrial  80  

    spheres  as  well  as  chemical  components  of  the  atmosphere”.  81  

    The Twain/Heinlein expression was   strongly   endorsed   by   the   late   E.   Lorenz  82  

    who   stated:   “Before   embarking   on   a   search   for   an   ideal   definition   (of   climate)  83  

    assuming   one   exists,   let   me   express   my   conviction   that   such   a   definition,   when  84  

    found  must  agree  in  spirit  with  the  statement,  “climate  is  what  you  expect”.”  [Lorenz, 85  

    1995].     He   then   proposed   several   definitions   based   on   dynamical   systems   theory  86  

    and  strange  attractors  (see  also  [Palmer, 2005]).      87  

    A variant on this, motivated by GCM modeling, was proposed by [Bryson, 1997] 88  

    (criticized by [Pielke, 1998]): “Climate is the thermodynamic/ hydrodynamic status of 89  

    the global boundary conditions that determine the concurrent array of weather patterns.” 90  

    He explains that whereas “weather forecasting is usually treated as an initial value 91  

    problem … climatology deals primarily with a boundary condition problem and the 92  

    patterns and climate devolving there from”. This definition could be paraphrased “for 93  

    given boundary conditions, the climate is what you expect”. This and similar views 94  

    provide the underpinnings for much of current climate prediction, including the recent 95  

    idea of “seamless forecasting” (e.g. [Palmer et al., 2008],   [Palmer, 2012]) in which 96  

    seasonal scale model validation is applied to climate scale predictions (for a recent 97  

    discussion, see [Pielke et al., 2012]). 98  

    There are two basic problems with the Twain/Heinlein dictum and its variants. 99  

  • Page 6

    The first is that they are based on an abstract weather - climate dichotomy, they are not 100  

    informed by empirical evidence. The glaring question of how long is “long” is either 101  

    decided subjectively or taken by fiat as the WMO’s “normal” 30 year period. The second 102  

    problem – that will be evident momentarily – is that it assumes that the climate is nothing 103  

    more than the long-term statistics of weather. If we accept the usual interpretation – that 104  

    we “expect” averages - then the dictum means that averaging weather over long enough 105  

    time periods converges to the climate. With regards to external forcings, one could argue 106  

    that this notion could still implicitly include the atmospheric response i.e. with averages 107  

    converging to slowly varying “responses”. However, it implausibly excludes the 108  

    appearance of any new “slow”, internal climate processes leading to fluctuations growing 109  

    rather than diminishing with scale. 110  

    To overcome this objection, one might adopt a more abstract interpretation of 111  

    what we “expect”. For example if the climate is defined as the probability distribution of 112  

    weather states, then all we “expect” is a random sample. However even this works only 113  

    inasmuch as it is possible to merge fast weather processes with slow climate processes 114  

    into a single process with a well defined probability distribution. While this may satisfy 115  

    the theoretician, it is unlikely to impress the layman. This is particularly true since to be 116  

    realistic we will see that one of the regimes in this composite model must have the 117  

    property that fluctuations converge whereas in the other, they diverge with scale. To use 118  

    the single term “climate” to encompass the two opposed regimes therefore seems at best 119  

    unhelpful and at worst misleadling. 120  

    The main purpose of this paper is show that the weather-climate dichotomy is 121  

    empirically untenable, that hiding between the two is a missing middle range spanning a 122  

  • Page 7

    factor of a thousand in scale (≈ 10 days to ≈ 30 -100 years) characterized by qualitatively 123  

    different dynamics. This new low frequency weather regime was dubbed “macroweather” 124  

    since it is a kind of large scale weather (not small scale climate) regime [Lovejoy and 125  

    Schertzer, 2013], it fundamentally clarifies the distinction between weather and climate, 126  

    the status and role of GCM models and the notion of climate predictability. 127  

    2.  The  variability  characterized  by  spectral  composites  128  

    Notwithstanding   the   existence   of   several   strong   periodicities   (notably  129  

    diurnal  and  annual),   the  atmosphere   is  highly  variable  over  huge  ranges  of   space-‐130  

    time  scales.     In  addition,   it  has   long  been  recognized  (e.g.   [Lovejoy and Schertzer, 131  

    1986],   [Wunsch, 2003])   that  even  at   the   longer  climate   scales,   the  contribution   to  132  

    the   variability   from   specific   frequencies   associated   with   specific   quasi   periodic  133  

    processes   is   small:   the   great  majority   of   the   variance   in   the   spectrum   is   from   the  134  

    continuous   background   “noise”.     Any   objectively   based   definition   of   weather   or  135  

    climate   must   therefore   start   from   a   clear   picture   of   the   atmosphere’s   temporal  136  

    variability  over  wide  scale  ranges.      137  

    The   first   -‐   and   still   most   ambitious   -‐   single   composite   spectrum   of  138  

    atmospheric  variability  [Mitchell, 1976],  ranged  from  hours  to  the  age  of  the  earth  (≈  139  

    10-‐4  to  1010  yrs),  fig.  1  shows  a  modern  version.    Given  the  rudimentary  quality  of  the  140  

    data   at   that   time,   Mitchell   admitted   that   his   composite   was   mostly   an   “educated  141  

    guess”   (indeed,   over   the   range   ≈   104   to   ≈   108   yrs,   [Shackleton and Imbrie, 1990]  142  

    empirically   found   that   Mitchell   had   underestimated   the   variation   of   the   spectral  143  

    density  by  a  factor  of  ≈  105!).    His  framework  reflected  the  prevailing  idea  that  there  144  

  • Page 8

    were   numerous   roughly   periodic   processes,   superposed   onto   a   continuous  145  

    background   spectrum  E(ω)  made   up   of   a   hierarchy   of   white   noise   processes   and  146  

    their  integrals  (i.e.  Ornstein-‐Uhlenbeck  processes  with  spectra  E(ω)  ≈  ω-‐β  with  β  =  0,  147  

    2   respectively  where  β   is   the   spectral   exponent:   the  negative   slope  on   the   log-‐log  148  

    plot   in   fig.   1).     The   spectral   spikes   were   therefore   superposed   on   a   spectrum  149  

    consisting   of   a   series   of   “shelves”   and   represented   distinct   physical   processes.    150  

    Mitchell  explained  his  idea  as  follows:  151  

     152  

    “As  we  scan  the  spectrum  from  the  short-‐wave  end  toward  the  longer  wave  153  

    regions,   at   each   point   where   we   pass   through   a   region   of   the   spectrum  154  

    corresponding   to   the   time  constant  of  a  process   that  adds  variance   to   the  climate,  155  

    the   amplitude   of   the   spectrum   increases   by   a   constant   increment   across   all  156  

    substantially   longer  wavelengths.     In   other  words,   each   stochastic   process   adds   a  157  

    shelf  to  the  spectrum  at  an  appropriate  wavelength”  [Mitchell, 1976].  158  

     159  

    By   the  early  1980’s,   following   the  explosion  of  scaling  (fractal)   ideas   it  was  160  

    realized   that   scale   invariance   was   a   very   general   symmetry   principle   often  161  

    respected   by   nonlinear   dynamics,   including   many   geophysical   processes   and  162  

    turbulence.    The  signature  of  a  scaling  process  is  a  power  law  spectrum,  linear  on  a  163  

    log-‐log  plot.    Although   in  order   to  accommodate   the  wide  range  of  scales,  Mitchell  164  

    had   found   it   “necessary   to   resort   to   logarithmic   coordinates”,   there   was   no  165  

    implication   that   the   underlying   processes   might   have   nontrivial   scaling   over   any  166  

    significant  range.    In  contrast,  scaling  symmetries,  were  explicitly  invoked  to  justify  167  

  • Page 9

    the   alternative   composite   picture   ([Lovejoy and Schertzer, 1984; Lovejoy and 168  

    Schertzer, 1986]  which  profited  from  early  ocean  and  ice  core  paleotemperatures.    169  

    These  analyses  already  clarified  the  following  points:  a)  the  distinction  between  the  170  

    variability   of   regional   and   global   scale   temperatures,   b)   that   there   was   a   scaling  171  

    range   for  global  averages  between  scales  of   the  order  of  10  yrs   (τc   in   the  notation  172  

    here)  up  to  ≈  40  -‐  50  kyrs  with  an  exponent  βc  ≈  1.8,  c)  that  a  scaling  regime  with  this  173  

    exponent   could   quantitatively   explain   the   magnitudes   of   the   temperature   swings  174  

    between  interglacials  (“ice  ages”):  the  “interglacial  window”  (see  below).  175  

    In   the   last   15   years   this   picture   has   been   supported   by   the   quite   similar  176  

    scaling  composites  of  [Pelletier, 1998]  and  [Huybers and Curry, 2006].    The  latter  in  177  

    particular   made   a   data   intensive   study   of   the   scaling   of   many   different   types   of  178  

    paleotemperatures  collectively  spanning   the  range  of  about  1  month   to  nearly  106  179  

    years.     In   addition,   even  without   producing   composites,   other   authors   shared   the  180  

    scaling   framework,   e.g.   [Koscielny-Bunde et al., 1998],   [Talkner and Weber, 2000],  181  

    [Ashkenazy et al., 2003; Rybski et al., 2008].    Their   results  are  qualitatively  very  182  

    similar  -‐  including  the  positions  of  the  scale  breaks;  the  main  innovations  are  a)  the  183  

    increased  precision  on   the  β   estimates  and  b)   the  basic  distinction  made  between  184  

    continental  and  oceanic  spectra   including  their  exponents.    We  could  also  mention  185  

    the  composite  of  [Fraedrich et al., 2009]  which  is  a  modest  adaptation  of  Mitchell’s:  186  

    it  innovates    by  introducing  a  single  scaling  regime  from  ≈  3  to  ≈  100  yrs.  187  

    Using   real   temperature   and   paleotemperature   data,   examples   showing   the  188  

    behaviors  in  the  three  different  regimes  are  graphically  illustrated  in  fig.  2.    Notice  189  

    that  in  the  weather  regime  (bottom)  the  temperature  seems  to  “wander”  up  or  down  190  

  • Page 10

    like  a  drunkard’s  walk  so  that  temperature  differences  typically  increase  over  longer  191  

    and  longer  periods.    Turning  to  the  middle  macroweather  series,  we  see  that  it  has  a  192  

    totally  different  appearance  with  successive  fluctuations  on  the  contrary  tending  to  193  

    cancel  each  other  out,   i.e.  with  decreases  followed  by  partially  cancelling  increases  194  

    (and   visa   versa).     At   first   sight,   this   vindicates   the   “climate   is  what   you   get”   idea  195  

    since  averages  over  longer  and  longer  periods  will  clearly  converge.    From  this,  we  196  

    anticipate  that  at  decadal  -‐  or  certainly  at  centennial  scales  -‐  that  we  will  see  at  most  197  

    smooth,  slow  variations.    However,  when  we  turn  to  the  century  resolution  climate  198  

    series  (top)  on  the  contrary,  we  once  again  see  weather-‐like  wandering.      199  

    Below  we  shall   see   that   the   fluctuations  over  a   time   interval  Δt   are   in  each  200  

    case  roughly  scaling  (power  laws)  of  the  form  ΔtH  so  that  the  sign  of  H  qualitatively  201  

    distinguishes   the   “wandering”   (H>0)   or   “cancelling”   (H

  • Page 11

    Starting  with  the  climate  regime,  numerous paleo temperature series (mostly from ice 213  

    and ocean cores) have been analyzed and there is broad agreement on their scaling nature 214  

    with spectral exponents estimated in the range βc ≈ 1.3 to 2.1 over range from hundreds to 215  

    tens of thousands of years,   [Lovejoy and Schertzer, 1986],   [Schmitt et al., 1995],  216  

    [Ditlevsen et al., 1996],  [Pelletier, 1998],  [Ashkenazy et al., 2003],  [Wunsch, 2003],  217  

    [Huybers and Curry, 2006],  [Blender et al., 2006] ,  [Lovejoy and Schertzer, 2012c]. 218  

    These analyses employed diverse techniques including spectra, difference and Haar 219  

    structure functions as well as Detrended Fluctuation Analysis so that the results are fairly 220  

    robust. In addition, as discussed below (fig. 3), further analyses from surface 221  

    temperatures, multiproxy reconstructions and 138 year long Twentieth Century reanalysis 222  

    (20CR), lend this further quantitative support. 223  

    Similarly,  in  the  macroweather  regime,  there  are  now  many  studies  finding  224  

    scaling   with   spectral   exponents   βmw  

  • Page 12

    Of   the   three   regimes,   the   only   one   where   the   idea   of   a   roughly   scaling  236  

    background   spectrum   is   still   somewhat   controversial   is   the   higher   frequency  237  

    weather  regime  (scales  <  τw  ≈  10  days).    To  understand  the  debate,  recall   that   the  238  

    classical   turbulence   theories   describing   the   statistical   variability   in   the   weather  239  

    regime  are  all  based  on  isotropic  scaling,  the  most  famous  being  the  Kolmogorov  k-‐240  

    5/3   spectrum   for   the   wind   (k   is   a   wavenumber).     However,   the   strong   vertical  241  

    atmospheric   stratification   prevents   isotropic   scaling   from   holding   over   any   scale  242  

    ranges   spanning   the   scale   thickness   of   the   atmosphere   (   ≈   10   km).     One   must  243  

    therefore   abandon   either   the   scaling   or   the   isotropy   assumption.     Following  244  

    Kraichnan’s   development   of   2-‐D   turbulence   and   Charney’s   extension   to   (still  245  

    essentially   2D)   “quasi   geostrophic”   turbulence,   the   usual   choice   was   to   retain  246  

    isotropy  and  to  divide  the  dynamics  into  2D  isotropic  (large  scale)  and  3D  isotropic  247  

    (small)  scale  regimes  ([Kraichnan, 1967],  [Charney, 1971]).    However,  starting  with  248  

    [Schertzer and Lovejoy, 1985],   a   growing   body   of   evidence   and   theory   has  249  

    supported   the   alterative   anisotropic   scaling   hypothesis.     Thanks   both   to   modern  250  

    empirical   evidence   (e.g.   the   review   [Lovejoy and Schertzer, 2010],   [Lovejoy and 251  

    Schertzer, 2013]  and  a  recent  massive  aircraft  study  [Pinel et al., 2012]),  but  also  to  252  

    theoretical   arguments   showing   that   the   governing   equations   are   symmetric   with  253  

    respect   to   anisotropic   scaling   symmetries   ([Schertzer et al., 2012]),   the   question  254  

    increasingly  has  been  settled   in   favor  of  anisotropic  scaling  (see  the  recent  debate  255  

    [Lovejoy et al., 2009],   [Lindborg et al., 2010a; Lindborg et al., 2010b],   [Lovejoy et 256  

    al., 2010],   [Schertzer et al., 2011],   [Yano, 2009],   ([Schertzer et al., 2012]).     The  257  

  • Page 13

    implications   of   this   anisotropic   spatial   scaling   for   the   temporal   statistics   are  258  

    discussed  in  [Radkevitch et al., 2008]  and  [Lovejoy and Schertzer, 2010].  259  

    A   review   of   diverse   evidence   from   reanalyses,   in   situ   and   remotely   sensed  260  

    data   ([Lovejoy and Schertzer, 2010],   [Lovejoy and Schertzer, 2013])   shows   that  261  

    for   wind,   temperature,   humidity,   pressure,   short   and   long   wave   radiances,   βw   is  262  

    commonly   in   the   range  1.5   -‐2   (certainly  >1,  hence  H>0).    The  existence  of   a  basic  263  

    transition   in   the   range   ≈  5   –   20   days   has   been   recognized   at   least   since   [Van der 264  

    Hoven, 1957]  who  noted  a  low  frequency  spectral  “bump”  at  around  5  days.    Later,  265  

    the   corresponding   features   in   the   temperature   and  pressure   spectra  were   termed  266  

    “synoptic  maxima”   by   [Kolesnikov and Monin, 1965]   and   [Panofsky, 1969].    More  267  

    recently,   in   the   same   spirit   as   Mitchell,   the   transition   has   been   modeled   (e.g.  268  

    [AchutaRao and Sperber, 2006])  as  an  Orenstein-‐Uhlenbeck  process  i.e.  with  βw  =  2,  269  

    βmw  =   0,   corresponding   to  Hw   =   1/2,  Hmw  =   -‐1/2,   although   as   can   be   seen   in   fig.   3  270  

    (discussed  below),  this  is  not  a  very  accurate  approximation  and  can  be  misleading.    271  

    Finally, [Vallis, 2010] proposed a (nonscaling) mechanism by suggesting that τw is 272  

    determined by the lifetimes of baroclinic instabilities. These were estimating by the 273  

    inverse Eady growth rate (τEady) which yielded τEady ≈ 4 days. However this result 274  

    requires a linearization of the equations about a hypothetical state having uniform shear 275  

    and uniform stratification across the entire troposphere. In comparison, the real  276  

    troposphere   has   highly   nonuniform   shears   and   stratifications,   its   variability   is   so  277  

    strong  that  it  is  characterized  by  anomalous  scaling  exponents  throughout  [Lovejoy 278  

    et al., 2007]     (including  H  ≈  0.75  for  the  horizontal  wind  in  the  vertical  direction).    279  

    Another difficulty with using τEady ≈ τw is that τEady is inversely proportional to the 280  

  • Page 14

    Coriolis parameter so that it diverges at the equator whereas the empirical τw is not very 281  

    sensitive to latitude. 282  

    Indeed,  a   seductive   feature  of   the   (anisotropic)  scaling   framework   is   that   it  283  

    fairly  accurately  predicts  the  weather  to  macroweather  transition  scale  τw  ≈  10  days.    284  

    The   argument   is   as   follows:   the   sun   provides   ≈   240  W/m2   of   heating   with   a   2%  285  

    efficiency  of  conversion  to  kinetic  energy  ([Monin, 1972]).    The  energy  is  distributed  286  

    reasonably   uniformly   over   the   troposphere,   leading   to   a   turbulent   energy   flux  287  

    density  (ε)  close  to  the  observed  global  value  ε  ≈  10-‐3  W/kg  ([Lovejoy and Schertzer, 288  

    2010];  this  is  the  flux  of  energy  from  large  to  small  scales).    The  model  predicts  that  289  

    this  turbulent  energy  flux  is  the  fundamental  driver  of  the  horizontal  dynamics  and  290  

    thus   that   planetary   structures   have   eddy-‐turnover   times   of   ≈   ε-‐1/3Le2/3   ≈   10   days  291  

    where  Le  =20000  km  is  the  largest  great  circle  distance  on  the  earth.    The  analogous  292  

    calculation   for   the   ocean   using   the   empirical   ocean   turbulent   flux   ε   ≈   10-‐8  W/kg,  293  

    yields   a   lifetime   of   ≈   1   yr   which   is   indeed   the   scale   separating   a   high   frequency  294  

    “ocean  weather”  (with  β  >1)  from  a  low  frequency  “macro-‐ocean  weather”  with  β  

  • Page 15

    structures  with   lifetimes  Δt;   at   scales  Δt>τw   they  depend  on   the   statistics  of  many  304  

    planetary  sized  structures.     In  addition,   the  basic  FIF  model  predicts   [Lovejoy and 305  

    Schertzer, 2012c]  macroweather  exponents   typically   in   the   range    0.2  <  βmw  <  0.6  306  

    (i.e.  -‐0.4  <  Hmw  <  -‐0.2).      307  

    4.  Real  space  fluctuations  and  analyses  308  

    In   spite   of   the   now   burgeoning   evidence   that   the   atmosphere’s   natural  309  

    variability   is  scaling  over  various  ranges,   the   idea  has  not  received  the  attention   it  310  

    deserves   and   at   least   over   decadal,   centennial   and   millennial   scales,   the   natural  311  

    variability   is   still   largely   identified   with   quasi-‐periodic   behaviours   which   –   when  312  

    present   -‐   have   the   advantage   of   being   predictable   (for   examples   of   periodicities  313  

    ranging   from   multidecadal   to   millennial   scales   see   [Delworth et al., 1993],  314  

    [Schlesinger and Ramankutty, 1994],  [Mann and Park, 1994],  [Mann et al., 1995],  315  

    [Bond et al., 1997],   [Isono et al., 2009]).     An   additional   reason   for   this   focus   on  316  

    quasi-‐periodic  behavior  is  that  whereas  spectra  are  ideal  for  understanding  periodic  317  

    processes,  they  are  not  optimal  for  scaling  processes.    For  these,  the  corresponding  318  

    real   space  analyses  are  more   straightforward   to   interpret;   this   is  particularly   true  319  

    when   comparing   spectra   from  different   data   types  with   different   resolutions.       In  320  

    this  section  we  show  how  this  works.  321  

     In  order  to  understand  the  qualitatively  different  behaviors  in  fig.  2,  consider  322  

    fluctuations  ΔT.    In  a  scaling  regime,  these  will  change  with  scale  as  ΔT  = ϕ  ΔtH  where  323  

    H is the fluctuation exponent (also called the “nonconservation” exponent; it is denoted 324  

    “H” in honor of Edwin Hurst but in general – unless the process is Gaussian - it is not the 325  

  • Page 16

    same as the Hurst exponent). ϕ is a controlling dynamical variable (e.g. a turbulent flux) 326  

    whose mean < ϕ > is independent of the lag Δt (i.e. independent of the time scale). The 327  

    behavior of the mean fluctuation is thus ≈   ΔtH so that if H>0, on average 328  

    fluctuations tend to grow with scale whereas if H0),  mean  335  

    absolute   differences   cannot   decrease   and   so   when   H

  • Page 17

    Once  estimated,  the  variation  of  the  fluctuations  with  scale  can  be  quantified  347  

    by   using   their   statistics;   the   qth   order   structure   function   Sq(Δt)   is   particularly  348  

    convenient:  349  

    Sq Δt( ) = ΔT Δt( )q (1) 350  

    where   “”   indicates   ensemble   averaging.     In   a   scaling   regime,   Sq(Δt)   is   a   power  351  

    law;   Sq(Δt)   ≈   Δtξ(q),   where   the   exponent   ξ(q)   =   qH   –   K(q)   where   K(1)   =   0.   Since  352  

    Gaussian   processes   have   K(q)   =   0,   K(q)   characterizes   the   strong   non   Gaussian  353  

    variability;  the  “intermittency”.    In  the  macroweather  regime  K(2)  is  typically  small  354  

    (≈0.01-‐0.03),  so   that   the  RMS  variation  S2(Δt)1/2  (denoted  simply  S(Δt)  below)  has  355  

    the  exponent  ξ(2)/2  ≈  ξ(1)  =  H.    In  the  climate  regime  this  intermittency  correction  356  

    is  a  bit  larger  [Schmitt et al., 1995]  but  the  error  in  using  this  approximation  (  ≈0.06)  357  

    will  be  neglected.    Note  that  since  the  spectrum  is  a  second  order  statistic,  we  have  358  

    the  useful  relationship  β  =  1+ξ(2)  =  1+2H-K(2).    When  K(2)  is  small,  β  ≈  1+2H  so  that  359  

    as  mentioned  earlier,  H>0,  H  1,  β0,  Hmw0).    The  transitions  are  at  τw  ≈  5  -‐  10  365  

    days  and  τc  ≈  30  -‐100  yrs:  note  that  in  the  recent  period τc  is  a  bit  lower  (≈  10  yrs)  -‐  366  

    see  the  1880-‐2008  surface  series  –  than  in  earlier  periods  (see  the  multiproxies  with  367  

    τc   ≈  30   -‐100  yrs),   this   is  because  over   the   last   century,   anthropogenic   forcing  has  368  

  • Page 18

    become  dominant  for  scales  greater  than  about  10  yrs,  see  [Lovejoy, 2013].    We  can  369  

    also  glimpse  a   fourth   low   frequency   “macro  climate”   regime   for   scales   larger   than  370  

    τmc  ≈  100  kyrs,  but  this  is  outside  our  scope.    b)  the  difference  between  the  local  and  371  

    global  fluctuations.    c)  the  “glacial/interglacial  window”  corresponding  to  overall  ±3  372  

    to  ±5  K  variations  (i.e.  S(Δt)  ≈  6,  10  K)  over  scales  with  half  periods  of  30  –  50  kyrs;  373  

    the  curve  must  pass  through  the  window  in  order  to  explain  the  glacial/interglacial  374  

    transitions.     Starting   at   τc   ≈   10   -‐   30   yrs,   one   can   plausibly   extrapolate   the   global  375  

    surface   and   20CR   700   mb   S(Δt)’s   using   H   =   0.4   (β   ≈1.8),   all   the   way   to   the  376  

    interglacial  window  (with  nearly  an  identical  S  as  in  [Lovejoy and Schertzer, 1986]).    377  

    Similarly,   the   local   temperatures   and   multiproxies   also   seem   to   follow   the   same  378  

    exponent   with   slightly   different   τc’s   and   seem   to   extrapolate   respectively   a   little  379  

    above  and  below  the  window.    380  

    These   statistics  may   seem   arcane   but   their   physical   interpretation   is   pretty  381  

    straightforward.     In   the   weather   regime,   larger   and   larger   fluctuations   “live”   for  382  

    longer   and   longer   times:   the   “eddy   turnover   time”.       At   any   given   time   scale,   the  383  

    fluctuations  are  dominated  by  structures  with  corresponding  spatial  scales  and  this  384  

    relationship  holds  up  to  structures  of  planetary  scales  with  lifetimes  ≈  10  days.    For  385  

    periods  longer  than  this,  the  statistics  are  dominated  by  averages  of  many  planetary  386  

    scale   structures,   and   these   fluctuations   tend   to   cancel   out:   for   example   large  387  

    temperature   increases   are   typically   followed   (and   partially   cancelled)   by  388  

    corresponding  decreases.  The  consequence  is  that  in  this  macroweather  regime,  the  389  

    average  fluctuations  diminish  as  the  time  scale  increases.    At  some  point  –  at  around  390  

    30  years  depending  on  geographic  location  and  time  (and  as  little  as  10  years  in  the  391  

  • Page 19

    recent   period   when   anthropogenic   variability   is   important),   these   weaker   and  392  

    weaker  fluctuations  -‐  whose  origin  is  in  weather  dynamics  -‐  become  dominated  by  393  

    increasingly   strong   lower   frequency   processes.       These   not   only   include   changing  394  

    external   solar,   volcanic   orbital   or   anthropogenic   “forcings”   –   but   quite   likely   also  395  

    new  and  increasingly  strong  slow  (internal)  climate  processes  -‐  or  by  a  combination  396  

    of  the  two:  forcings  with  feedbacks.    A  relevant  example  of  a  slow  dynamical  process  397  

    that  is  not  currently  fully  incorporated  into  GCM’s  is  land-‐ice  dynamics.    The  overall  398  

    effect   is   that   in   the   resulting   climate   regime,   fluctuations   tend   to   grow  again  with  399  

    scale   in   an   “unstable”  manner,   very   similar   to   the  way   they   grow   in   the  weather  400  

    regime.  401  

    5.  Implications  for  climate  modelling,  prediction  402  

    Numerical  weather  models  and  reanalyses  are  qualitatively  in  good  agreement  403  

    with  the  weather  /  macroweather  picture  described  above,  although  there  are  still  404  

    some   quantitative   discrepancies   in   the   values   of   the   exponents,   possibly   due   the  405  

    hydrostatic   approximation   and   other   numerical   issues   ([Stolle et al., 2009],  406  

    [Lovejoy and Schertzer, 2011]).    However,  climate  models   (GCM’s)  are  essentially  407  

    weather  models  with  various  additional  couplings  (with  ocean,  carbon  cycle,   land-‐408  

    use,  sea  ice  and  other  models).     It   is  therefore  not  surprising  that  control  runs  (i.e.  409  

    with  no  “climate  forcings”)  generate  macroweather  (with  βmw  ≈  0.2,  Hmw  ≈  -‐0.4),  and  410  

    this  apparently  out  to  the  extreme  low  frequency  limit  of  the  models  (see  the  review  411  

    and  analyses  in  [Lovejoy et al., 2012]  as  well  as  [Blender et al., 2006],  [Rybski et al., 412  

    2008]).    413  

  • Page 20

    Avoiding   anthropogenic   effects   by   considering   the   pre-‐1900   epoch,   for   GCM  414  

    climate  models,  the  key  question  is  whether  solar,  volcanic,  orbital  or  other  climate  415  

    forcings  are  sufficient  to  arrest  the  H  0   regime  with   sufficiently   strong   centennial,  millennial   variability   to  417  

    account  for  the  background  variability  out  to  glacial-‐  interglacial  scales.    Analysis  of  418  

    several   last   millennium   simulations   has   found   that   at   the   moment,   their   low  419  

    frequency  variabilities  are  too  weak  [Lovejoy et al., 2012].  420  

    To  understand  this  weak  variability,  one  can  examine  the  scale  dependence  of  421  

    fluctuations   in   the   radiative   forcings   (ΔRF)   of   several   solar   and   volcanic  422  

    reconstructions;  they  are  generally  scaling  with  ΔRF  ≈  AΔtHR  [Lovejoy and Schertzer, 423  

    2012a].    If  HR  ≈HT  ≈  0.4,  then  scale  independent  amplification  /  feedback  mechanisms  424  

    would  suffice.    However  it  was  often  found  that  HR  ≈  -‐0.3  implying  that  the  forcings  425  

    become   weaker   with   scale   -‐   even   though   the   response   grows   with   scale.     This  426  

    suggests  the  need  to  introduce  new  slow    mechanisms  of  internal  variability.    Such  427  

    mechanisms   must   have   broad   spectra;   this   suggests   that   their   dynamics   involve  428  

    nonlinearly   interacting   spatial   degrees   of   freedom   such   as   the   land-‐ice   dynamics  429  

    mentioned  above.  430  

    Whatever  the  ultimate  source  of  the  growing  fluctuations  in  the  H  >  0  climate  431  

    regime,  a  careful  and  complete  characterization  of  the  scaling  in  space  as  well  as  in  432  

    time  (including  possible  space-‐time  anisotropies)  allows  for  new  stochastic  methods  433  

    for  predicting  the  climate.    The   idea   is   to  exploit   the  particularly   low  variability  of  434  

    the  averages  at  scale  τc.    Since  τc  ≈  30  yrs  -‐  i.e.  the  conventional  but  ad  hoc  “climate  435  

    normal”  period   -‐   this  not  only   justifies   the  normal  but  allows  averages  of   relevant  436  

  • Page 21

    variables  over  it  to  define  “climate  states”  and  the  changes  at  scales  Δt>τc  to  define  437  

    climate   change   (again,   in   the   recent   period,   this   defines   the   scale   at   which  438  

    anthropogenic   variability   starts   to   dominate   natural   variability).     Even   without  439  

    resolving  the  question  of  the  dominant  climate  forcing  and  slow  internal  feedbacks,  440  

    one  could  use  the  statistical  properties  of  the  climate  states  -‐  the  system’s  “memory”  441  

    implicit  in  the  long  range  statistical  correlations  –  combined  with  the  growing  data  442  

    on  past  climate  states  in  order  to  make  stochastic  climate  forecasts.  443  

    Another  attractive  application  of  this  scaling  picture  is  that  by  quantifying  the  444  

    natural  variability  as  a   function  of   space-‐time  scales,   it  opens  up   the  possibility  of  445  

    convincingly   distinguish   natural   and   anthropogenic   variability.     This   is   possible  446  

    because   the   stochastic   scaling   framework   allows   one   to   statistically   test   specific  447  

    hypotheses  about  the  probability  that  the  atmosphere  would  naturally  behave  in  the  448  

    way   that   is   observed,   i.e.   to   formulate   rigorous   statistical   tests   of   any   trends   or  449  

    events  against  the  null  hypothesis.    Only  if  the  probabilities  are  low  enough  should  450  

    the   hypothesis   that   the   observed   changes   are   natural   in   origin   be   rejected.     For  451  

    example,  using  CO2  as  a  surrogate  for  all  anthropogenic  effects  and  taking  into  both  452  

    long  range  statistical  dependencies  and  extreme  “fat  tailed”  probabilities,  [Lovejoy, 453  

    2013]   found   that   the   probability   of   the   global   warming   since   1880   being   due   to  454  

    natural   variability   can   be   rejected   with   more   than   99%   confidence.     The   scaling  455  

    flucutation   approach   thus   allows   quantitative   (and   hence   convincing)   answers   to  456  

    questions  such  as:  how  can  the  earth  have  prolonged  periods  of  cooling  in  the  midst  457  

    of   anthropogenic   warming;   or   was   this   winter’s   record   mild   temperature   really  458  

    evidence  for  anthropogenic  influence?    Finally,  the  systematic  comparison  of  model  459  

  • Page 22

    and  natural  variability   in   the  preindustrial  era   is   the  best  way  to   fully  address   the  460  

    issue  of  “model  uncertainty”,  to  assess  the  extent  by  which  the  models  are  missing  461  

    important  slow  processes.    462  

    6.  Conclusions  463  

    Contrary to [Bryson, 1997], we have argued that the climate is not accurately 464  

    viewed as the statistics of fundamentally fast weather dynamics that are constrained by 465  

    quasi fixed boundary conditions. The empirically substantiated picture is rather one of 466  

    “unstable”, “wandering”, high frequency weather processes (i.e. H>0) tending - at scales 467  

    beyond 10 days or so and primarily due to the quenching of spatial degrees of freedom 468  

    (intermediate frequency, low variability) to macroweather processes. These appear to be 469  

    stable because positive and negative fluctuations tend to cancel out (H0) and only emerge from 471  

    macroweather at even lower frequencies, due to new slow internal climate processes 472  

    coupled with external forcings (including in the recent period, anthropogenic forcings). 473  

    These processes presumably include various nonlinear couplings with the fields that 474  

    Bryson considered to be no more than “boundary conditions” so that “rather than 475  

    ‘boundaries’ these become interactive mediums” [Pielke, 1998]. The 476  

    theoretical/mathematical picture underpinning GCM based prediction beyond ≈ 30 years 477  

    should thus be re-examined. 478  

    Yet, whatever the cause, it is an empirical fact that the emergent synergy of new 479  

    processes yields fluctuations that on average again grow with scale in at least a roughly 480  

    scaling manner and become dominant typically on time scales of 30 -100 years 481  

    (somewhat less in the recent period) up to ≈ 100 kyrs. 482  

  • Page 23

    Looked  at  another  way,   if   the  climate   really  was  what  you  expected,   then  –  483  

    since  one  usually  expects  averages  -‐  predicting  the  climate  would  be  the  relatively  484  

    simple  matter  of  determining   the   fixed   climate  normal.    On   the   contrary,  we  have  485  

    argued   that   from   the   stochastic   point   of   view   -‐   and   notwithstanding   the   vastly  486  

    different   time   scales   -‐   that   predicting   natural   climate   change   is   very   much   like  487  

    predicting   the   weather.     This   is   because   the   climate   at   any   time   or   place   is   the  488  

    consequence   of   climate   changes   that   are   (qualitatively   and   quantitatively)  489  

    unexpected  in  very  much  the  same  way  that  the  weather  is  unexpected.  490  

    At  a  subjective   level,   from  experience  over   the  years,  we  all  grow  to  expect  491  

    certain   stable   patterns   of   macroweather   (complicated   by   seasonal,   and   now   by  492  

    anthropogenic   effects,   but  nevertheless   recognizable   from  year   to   year)   so   that   in  493  

    daily  discourse  we  may  say  “macroweather  is  what  you  expect,  the  weather  is  what  494  

    you   get”.     However   over   generational   scales   -‐   periods   of   30   years   -‐   the  495  

    macroweather  we  are  accustomed  to  evolves  due  to  climate  change.    Speaking  to  our  496  

    children   and   grandchildren,   the   appropriate   dictum   would   therefore   be  497  

    “macroweather  is  what  you  expect,  the  climate  is  what  you  get”.  498  

    6.  Acknowledgements:  499  We  would   like   to   thank   Roger   Pielke,   Gavin   Schmidt,   and   Adrian   Tuck   for  500  

    useful  comments.  501  

    502  

  • Page 24

    7.  References  503  

    AchutaRao, K., and K. R. Sperber (2006), ENSO simulation in coupled ocean-504  atmosphere models: are the current models better? , Climate Dynamics, 27, 1–15 505  doi: 10.1007/s00382-006-0119-7. 506  

    Ashkenazy, Y., D. Baker, H. Gildor, and S. Havlin (2003), Nonlinearity and 507  multifractality of climate change in the past 420,000 years, Geophys. Res. Lett., 30, 508  2146 doi: 10.1029/2003GL018099. 509  

    Blender, R., K. Fraedrich, and B. Hunt (2006), Millennial climate variability: GCM‐510  simulation and Greenland ice cores, Geophys. Res. Lett., , 33, , L04710 doi: 511  doi:10.1029/2005GL024919. 512  

    Bond, G., W. Showers, M. Cheseby, R. Lotti, P. Almasi, P. deMenocal, P. Priori, H. 513  Cullen, I. Hajdes, and G. Bonani (1997), A pervasive millennial-scale climate cycle 514  in the North Atlantic: The Holocene and late glacial record, Science 278, 1257-515  1266. 516  

    Bryson, R. A. (1997), The Paradigm of Climatology: An Essay, Bull. Amer. Meteor. Soc. , 517  78, 450-456. 518  

    Charney, J. G. (1971), Geostrophic Turbulence, J. Atmos. Sci, 28, 1087. 519  Committee on Radiative Forcing Effects on Climate, N. R. C. (2005), Radiative Forcing 520  

    of Climate Change: Expanding the Concept and Addressing Uncertainties, 224 pp., 521  National Acad. press. 522  

    Delworth, T., S. Manabe, and R. J. Stoufer (1993), Interdecadal variations of the 523  thermocline ciruculation in a coupled ocean-atmosphere model, J. of Climate, 6, 524  1993-2011. 525  

    Ditlevsen, P. D., H. Svensmark, and S. Johson (1996), Contrasting atmospheric and 526  climate dynamics of the last-glacial and Holocene periods, Nature, 379, 810-812. 527  

    Eichner, J. F., E. Koscielny-Bunde, A. Bunde, S. Havlin, and H.-J. Schellnhuber (2003), 528  Power-law persistance and trends in the atmosphere: A detailed studey of long 529  temperature records, Phys. Rev. E, 68, 046133-046131-046135 doi: 530  10.1103/PhysRevE.68.046133. 531  

    Fraedrich, K., and K. Blender (2003), Scaling of Atmosphere and Ocean Temperature 532  Correlations in Observations and Climate Models, Phys. Rev. Lett., 90, 108501-533  108504 534  

    Fraedrich, K., R. Blender, and X. Zhu (2009), Continuum Climate Variability: Long-535  Term Memory, Scaling, and 1/f-Noise, , International Journal of Modern Physics B, 536  23, 5403-5416. 537  

    Heinlein, R. A. (1973), Time Enough for Love, 605 pp., G. P. Putnam's Sons, New York. 538  Huang, S. (2004), Merging Information from Different Resources for New Insights into 539  

    Climate Change in the Past and Future, Geophys.Res, Lett. , 31, L13205 doi: doi : 540  10.1029/2004 GL019781. 541  

    Huschke, R. E. (Ed.) (1959), Glossary of Meteorology, 638 pp. pp. 542  Huybers, P., and W. Curry (2006), Links between annual, Milankovitch and continuum 543  

    temperature variability, Nature, 441, 329-332 doi: 10.1038/nature04745. 544  

  • Page 25

    Isono, D., M. Yamamoto, T. Irino, T. Oba, M. Murayama, T. Nakamura, and H. 545  Kawahata ( 2009), The 1500-year climate oscillation in the midlatitude North 546  Pacific during the Holocene, Geology 37(591-594). 547  

    Kantelhardt, J. W., S. A. Zscchegner, K. Koscielny-Bunde, S. Havlin, A. Bunde, and H. 548  E. Stanley (2002), Multifractal detrended fluctuation analysis of nonstationary time 549  series, Physica A, 316, 87-114. 550  

    Kolesnikov, V. N., and A. S. Monin (1965), Spectra of meteorological field fluctuations, 551  Izvestiya, Atmospheric and Oceanic Physics, 1, 653-669. 552  

    Koscielny-Bunde, E., A. Bunde, S. Havlin, H. E. Roman, Y. Goldreich, and H. J. 553  Schellnhuber (1998), Indication of a universal persistence law governing 554  atmospheric variability, Phys. Rev. Lett. , 81, 729. 555  

    Kraichnan, R. H. (1967), Inertial ranges in two-dimensional turbulence, Physics of Fluids, 556  10, 1417-1423. 557  

    Lamb, H. H. (1972), Climate: Past, Present, and Future. Vol. 1, Fundamentals and 558  Climate Now, Methuen and Co. 559  

    Lanfredi, M., T. Simoniello, V. Cuomo, and M. Macchiato (2009), Discriminating low 560  frequency components from long range persistent fluctuations in daily atmospheric 561  temperature variability, Atmos. Chem. Phys., 9, 4537–4544. 562  

    Lennartz, S., and A. Bunde (2009), Trend evaluation in records with long term memory: 563  Application to global warming, Geophys. Res. Lett., 36, L16706 doi: 564  doi:10.1029/2009GL039516. 565  

    Lindborg, E., K. K. Tung, G. D. Nastrom, J. Y. N. Cho, and K. S. Gage (2010a), 566  Comment on “Reinterpreting aircraft measurement in anisotropic scaling turbulence” 567  by Lovejoy et al., , Atmos. Chem. Phys., , 10, 1401–1402. 568  

    Lindborg, E., K. K. Tung, G. D. Nastrom, J. Y. N. Cho, and K. S. Gage (2010b), 569  Interactive comment on “Comment on “Reinterpreting aircraft measurements in 570  anisotropic scaling turbulence" by Lovejoy et al. (2009)”, Atmos. Chem. Phys. 571  Discuss., 9 C9797–C9798. 572  

    Ljundqvist, F. C. (2010), A new reconstruction of temperature variability in the extra - 573  tropical Northern Hemisphere during the last two millennia, Geografiska Annaler: 574  Physical Geography, 92 A(3), 339 - 351 doi: DOI : 10.1111/j .1468 - 575  0459.2010 .00399.x. 576  

    Lorenz, E. N. (1995), Climate is what you expect, edited, p. 55pp, 577  aps4.mit.edu/research/Lorenz/publications.htm, (available,16 May, 2012). 578  

    Lovejoy, S. (2013), Anthropogenic warming quantified by scaling fluctuation analysis 579  Nature Clim Change, (submitted). 580  

    Lovejoy, S., and D. Schertzer (1984), 40 000 years of scaling in climatological 581  temperatures, Meteor. Sci. and Tech., 1, 51-54. 582  

    Lovejoy, S., and D. Schertzer (1986), Scale invariance in climatological temperatures and 583  the spectral plateau, Annales Geophysicae, 4B, 401-410. 584  

    Lovejoy, S., and D. Schertzer (2010), Towards a new synthesis for atmospheric 585  dynamics: space-time cascades, Atmos. Res., 96, pp. 1-52 doi: doi: 586  10.1016/j.atmosres.2010.01.004. 587  

    Lovejoy, S., and D. Schertzer (2011), Space-time cascades and the scaling of ECMWF 588  reanalyses: fluxes and fields, J. Geophys. Res., 116 doi: 10.1029/2011JD015654. 589  

  • Page 26

    Lovejoy, S., and D. Schertzer (2012a), Stochastic and scaling climate sensitivities: solar, 590  volcanic and orbital forcings, Geophys. Res. Lett. , 39, L11702 doi: 591  doi:10.1029/2012GL051871. 592  

    Lovejoy, S., and D. Schertzer (2012b), Haar wavelets, fluctuations and structure 593  functions: convenient choices for geophysics, Nonlinear Proc. Geophys. , 19, 1-14 594  doi: 10.5194/npg-19-1-2012. 595  

    Lovejoy, S., and D. Schertzer (2012c), Low frequency weather and the emergence of the 596  Climate, in Extreme Events and Natural Hazards: The Complexity Perspective, 597  edited by A. S. Sharma, A. Bunde, D. Baker and V. P. Dimri, pp. 231-254, AGU 598  monographs. 599  

    Lovejoy, S., and D. Schertzer (2013), The Weather and Climate: Emergent Laws and 600  Multifractal Cascades, 480 pp., Cambridge University Press, Cambridge. 601  

    Lovejoy, S., D. Schertzer, and A. F. Tuck (2010), Why anisotropic turbulence matters: 602  another reply to E. Lindborg, Atmos. Chem and Physics Disc., 10, C4689-C4697. 603  

    Lovejoy, S., D. Schertzer, and D. Varon (2012), Do GCM’s predict the climate…. or 604  macroweather?, Earth Syst. Dynam. Discuss., 3, , 1259-1286 doi: 10.5194/esdd-3-605  1259-2012. 606  

    Lovejoy, S., A. F. Tuck, S. J. Hovde, and D. Schertzer (2007), Is isotropic turbulence 607  relevant in the atmosphere?, Geophys. Res. Lett., doi:10.1029/2007GL029359, 608  L14802. 609  

    Lovejoy, S., A. F. Tuck, D. Schertzer, and S. J. Hovde (2009), Reinterpreting aircraft 610  measurements in anisotropic scaling turbulence, Atmos. Chem. and Phys. , 9, 1-19. 611  

    Mann, M. E., and J. Park (1994), Global scale modes of surface temperature variaiblity 612  on interannual to century timescales, J. Geophys. Resear., 99, 819-825. 613  

    Mann, M. E., J. Park, and R. S. Bradley (1995), Global interdecadal and century scale 614  climate oscillations duering the past five centuries, Nature, 378, 268-270. 615  

    Mitchell, J. M. (1976), An overview of climatic variability and its causal mechanisms, 616  Quaternary Res., 6, 481-493. 617  

    Moberg, A., D. M. Sonnechkin, K. Holmgren, and N. M. Datsenko (2005), Highly 618  variable Northern Hemisphere temperatures reconstructed from low- and high - 619  resolution proxy data, Nature, 433(7026), 613-617. 620  

    Monetti, R. A., S. Havlin, and A. Bunde (2003), Long-term persistance in the sea surface 621  temeprature fluctuations, Physica A, 320, 581-589. 622  

    Monin, A. S. (1972), Weather forecasting as a problem in physics, MIT press, Boston Ma. 623  Palmer, T. (2005), Global warming in a nonlinear climate – Can we be sure? , 624  

    Europhysics news March/April 2005, 42-46 doi: doi.org/10.1051/epn:2005202. 625  Palmer, T. N. (2012), Towards the probabilistic Earth-system simulator: a vision for the 626  

    future of climate and weather prediction, Q.J.R. Meteorol. Soc., in press. 627  Palmer, T. N., F. J. Doblas-Reyes, A. Weisheimer, and M. J. Rodwell (2008), Toward 628  

    Seamless Prediction: Calibration of Climate Change Projections Using Seasonal 629  Forecasts. , Bull. Amer. Meteor. Soc., , 89, 459–470 doi: doi.org/10.1175/BAMS-630  89-4-459. 631  

    Panofsky, H. A. (1969), The spectrum of temperature, J. of Radio Science, 4, 1101-1109. 632  Pelletier, J., D. (1998), The power spectral density of atmospheric temperature from 633  

    scales of 10**-2 to 10**6 yr, , EPSL, 158, 157-164. 634  

  • Page 27

    Peng, C.-K., S. V. Buldyrev, S. Havlin, M. Simons, H. E. Stanley, and A. L. Goldberger 635  (1994), Mosaic organisation of DNA nucleotides, Phys. Rev. E, 49, 1685-1689. 636  

    Pielke, R. (1998), Climate prediction as an initial value problem, Bull. of the Amer. 637  Meteor. Soc., 79, 2743-2746. 638  

    Pielke, R. A. S., R. Wilby, D. Niyogi, F. Hossain, K. Dairuku, J. Adegoke, G. Kallos, T. 639  Seastedt, and K. Suding (2012), Dealing with Complexity and Extreme Events 640  Using a Bottom-up, Resource-based Vulnerability Perspective, in Complexity and 641  Extreme Events in Geosciences, edited by A. S. Sharma, A. Bunde, D. Baker and V. 642  P. Dimri, AGU. 643  

    Pinel, J., S. Lovejoy, D. Schertzer, and A. F. Tuck (2012), Joint horizontal - vertical 644  anisotropic scaling, isobaric and isoheight wind statistics from aircraft data, 645  Geophys. Res. Lett., 39, L11803 doi: 10.1029/2012GL051698. 646  

    Radkevitch, A., S. Lovejoy, K. B. Strawbridge, D. Schertzer, and M. Lilley (2008), 647  Scaling turbulent atmospheric stratification, Part III: empIrical study of Space-time 648  stratification of passive scalars using lidar data, Quart. J. Roy. Meteor. Soc., DOI: 649  10.1002/qj.1203. 650  

    Rybski, D., A. Bunde, and H. von Storch (2008), Long-term memory in 1000- year 651  simulated temperature records, J. Geophys. Resear., 113, D02106-02101, D02106-652  02109 doi: doi:10.1029/2007JD008568. 653  

    Schertzer, D., and S. Lovejoy (1985), The dimension and intermittency of atmospheric 654  dynamics, in Turbulent Shear Flow 4, edited by B. Launder, pp. 7-33, Springer-655  Verlag. 656  

    Schertzer, D., and S. Lovejoy (1987), Physical modeling and Analysis of Rain and 657  Clouds by Anisotropic Scaling of Multiplicative Processes, Journal of Geophysical 658  Research, 92, 9693-9714. 659  

    Schertzer, D., I. Tchiguirinskaia, S. Lovejoy, and A. F. Tuck (2011), Quasi-geostrophic 660  turbulence and generalized scale invariance, a theoretical reply to Lindborg, Atmos. 661  Chem. and Physics Discussion, 11, 3301-3320. 662  

    Schertzer, D., I. Tchiguirinskaia, S. Lovejoy, and A. F. Tuck (2012), Quasi-geostrophic 663  turbulence and generalized scale invariance, a theoretical reply, Atmos. Chem. 664  Phys., 12, 327-336 doi: 10.5194/acp-12-327-2012. 665  

    Schlesinger, M. E., and N. Ramankutty (1994), An Oscillation in the Global Climate 666  System of Period 65-70 Years, Nature, 367, 723-726. 667  

    Schmitt, F., S. Lovejoy, and D. Schertzer (1995), Multifractal analysis of the Greenland 668  Ice-core project climate data., Geophys. Res. Lett, 22, 1689-1692. 669  

    Shackleton, N. J., and J. Imbrie (1990), The δ18O spectrum of oceanic deep water over a 670  five-decade band, Climatic Change, 16, 217-230. 671  

    Stolle, J., S. Lovejoy, and D. Schertzer (2009), The stochastic cascade structure of 672  deterministic numerical models of the atmosphere, Nonlin. Proc. in Geophys., 16, 673  1–15. 674  

    Talkner, P., and R. O. Weber (2000), Power spectrum and detrended fluctuation analysis: 675  Application to daily temperatures, Phys. Rev. E, 62, 150-160. 676  

    Vallis, G. (2010), Mechanisms of climate variaiblity from years to decades, in Stochstic 677  Physics and Climate Modelliing, edited by P. W. T. Palmer, pp. 1-34, Cambridge 678  University Press, Cambridge. 679  

  • Page 28

    Van der Hoven, I. (1957), Power spectrum of horizontal wind speed in the frequency 680  range from .0007 to 900 cycles per hour, Journal of Meteorology, 14, 160-164. 681  

    Wunsch, C. (2003), The spectral energy description of climate change including the 100 682  ky energy, Climate Dynamics, 20, 353-363. 683  

    Yano, J. (2009), Interactive comment on “Reinterpreting aircraft measurements in 684  anisotropic scaling turbulence” by S. Lovejoy et al., Atmos. Chem. Phys. Discuss., , 685  9, S162–S166 doi: http://www.atmos-chem-phys-discuss.net/9/S162/2009/. 686  

    687   688  

    689  

  • Page 29

    Figures and Captions: 690  

    691  

     692  

    Fig. 1: A composite temperature spectrum. Left: the GRIP (Summit) ice core δ18O (a 693  

    temperature proxy, low resolution (55 cm in depth back 240 kyrs) along with the first 91 694  

    kyrs at high resolution. Right: the spectrum of the (mean) 75oN of the 20th Century 695  

    Reanalysis (20CR) temperature spectrum, at 6 hour resolution, from 1871-2008, at 700 696  

    mb. The overlap (from 10 – 138 yr scales) is used for calibrating the former (moving 697  

    them vertically on the log-log plot). All spectra are averaged over logarithmically spaced 698  

    bins, ten per order of magnitude in frequency. Three regimes are shown corresponding to 699  

    the weather regime with βw= 2 (the diurnal variation and subharmonic at 12 hours are 700  

    visible at the extreme right). The central macroweather “plateau” is shown along with the 701  

    theoretically predicted βmw = 0.2 - 0.4 regime. Finally, at longer time scales (the left), a 702  

    MacroweatherClimate Weather

    m

  • Page 30

    new scaling climate regime with exponent βc ≈1.4 continues to about 100 kyrs. 703  

    Reproduced from [Lovejoy and Schertzer, 2012c]. 704  

     705  

     706  

    Fig.  2:  Dynamics  and  types  of  scaling  variability:    A  visual  comparison  displaying  707  

    representative   temperature   series   from  weather,  macroweather   and   climate   (H   ≈  708  

    0.4,  -‐0.4,  0.4,  bottom  to  top  respectively).  To  make  the  comparison  as  fair  as  possible,  709  

    in  each  case,   the  sample   is  720  points   long  and  each  series  has   its  mean  removed  710  

    and  is  normalized  by  its  standard  deviation  (4.49 K, 2.59 K, 1.39 K, respectively), the 711  

    two upper series have been displaced in the vertical by four units for clarity. The  712  

    resolutions  are  1  hour,  20  days  and  100  yrs,  the  data  are  from  a  weather  station  in  713  

    Lander  Wyoming   (detrended   daily,   annually),   the   20th   Century   reanalysis   and   the  714  

    100 200 300 400 500 600 700

    - 2

    2

    4

    6

    8

    10

    1 Century,Vostok,20-92kyr BP

    20 days,75oN,100 oW,1967 - 2008

    1 hour,Lander, 10 Feb.-12 March, 2005

    0

    T/

    t

  • Page 31

    Vostok   Antarctic   station. Note the similarity between the type of variability in the 715  

    weather and climate regimes (reflected in their scaling exponents although the H 716  

    exponent is only a partial characterization). 717  

       718  

  • Page 32

     719  

    Fig.  3:  Empirical  RMS  temperature  fluctuations  (S(Δt)),  local  scale  analyses:  On  720  

    the   left   top   we   show   grid   point   scale   (2ox2o)   daily   scale   fluctuations   globally  721  

    averaged   along   with   reference   slope   ξ(2)/2   =   -‐0.4   ≈   H   (20CR,   700   mb).     For 722  

    comparison, the results for 50 simulations of Orenstein-Uhlenbeck (OU) processes are 723  

    also given using simulations with a characteristic time of 3 days. The theoretical 724  

    asymptotic slopes (0.5, -0.5) are added to show their convergence to theory. Just  below  725  

    this,  we  show  the  monthly  NOAA CDC Sea Surface Temperature curve (5o resolution, 726  

    from 1900-2000); the transition from ξ(2)/2  ≈  0.4  to  ≈  -‐0.2  occurs  at  τow  ≈  1  yr.    On  the  727  

    lower  left,  we  see  at  daily  resolution,  the  corresponding  globally  averaged  structure  728  

    function.      729  

    Globally  averaged  series:    The  same  20CR  data  but  globally  averaged  (brown),  The  730  

    0.5

    0.5- 0

  • Page 33

    average  of  the  three  in  situ  surface  series  (NOAA  NCDC,  NASA  GISS,  CRU,  red)  as  well  731  

    as   the   average   of   three   post   2003  multiproxy   structure   functions;   [Huang, 2004],  732  

    [Ljundqvist, 2010; Moberg et al., 2005],  (see  [Lovejoy and Schertzer, 2012c]).      733  

    Paleotemperatures:   At   the   right  we   show   analysis   of   the   EPICA   Antarctic   series  734  

    interpolated   to  50  year   resolution  over  ≈  800  kyrs.    Also   shown   is   the   interglacial  735  

    “window”.  736  

    The reference slopes are ξ(2)/2 = -0.4, -0.2 or +0.4; these correspond to spectral 737  

    exponents β = 1+ξ(2) = 0.2, 0.6, 1.8 respectively; a flat curve in the above corresponds to 738  

    β = 1. 739  

     740  

  • Page 34

    Figure  Captions:  741  

    Fig. 1: A composite temperature spectrum. Left: the GRIP (Summit) ice core δ18O (a 742  

    temperature proxy, low resolution (55 cm in depth back 240 kyrs) along with the first 91 743  

    kyrs at high resolution. Right: the spectrum of the (mean) 75oN of the 20th Century 744  

    Reanalysis (20CR) temperature spectrum, at 6 hour resolution, from 1871-2008, at 700 745  

    mb. The overlap (from 10 – 138 yr scales) is used for calibrating the former (moving 746  

    them vertically on the log-log plot). All spectra are averaged over logarithmically spaced 747  

    bins, ten per order of magnitude in frequency. Three regimes are shown corresponding to 748  

    the weather regime with βw= 2 (the diurnal variation and subharmonic at 12 hours are 749  

    visible at the extreme right). The central macroweather “plateau” is shown along with the 750  

    theoretically predicted βmw = 0.2 - 0.4 regime. Finally, at longer time scales (the left), a 751  

    new scaling climate regime with exponent βc ≈1.4 continues to about 100 kyrs. 752  

    Reproduced from [Lovejoy and Schertzer, 2012c]. 753  

    Fig.  2:  Dynamics  and  types  of  scaling  variability:    A  visual  comparison  displaying  754  

    representative   temperature   series   from  weather,  macroweather   and   climate   (H   ≈  755  

    0.4,  -‐0.4,  0.4,  bottom  to  top  respectively).  To  make  the  comparison  as  fair  as  possible,  756  

    in  each  case,   the  sample   is  720  points   long  and  each  series  has   its  mean  removed  757  

    and  is  normalized  by  its  standard  deviation  (4.49 K, 2.59 K, 1.39 K, respectively), the 758  

    two upper series have been displaced in the vertical by four units for clarity. The  759  

    resolutions  are  1  hour,  20  days  and  100  yrs,  the  data  are  from  a  weather  station  in  760  

    Lander  Wyoming   (detrended   daily,   annually),   the   20th   Century   reanalysis   and   the  761  

    Vostok   Antarctic   station. Note the similarity between the type of variability in the 762  

  • Page 35

    weather and climate regimes (reflected in their scaling exponents although the H 763  

    exponent is only a partial characterization). 764  

    Fig.  3:  Empirical  RMS  temperature  fluctuations  (S(Δt)),  local  scale  analyses:  On  765  

    the   left   top   we   show   grid   point   scale   (2ox2o)   daily   scale   fluctuations   globally  766  

    averaged   along   with   reference   slope   ξ(2)/2   =   -‐0.4   ≈   H   (20CR,   700   mb).     For 767  

    comparison, the results for 50 simulations of Orenstein-Uhlenbeck (OU) processes are 768  

    also given using simulations with a characteristic time of 3 days. The theoretical 769  

    asymptotic slopes (0.5, -0.5) are added to show their convergence to theory. Just  below  770  

    this,  we  show  the  monthly  NOAA CDC Sea Surface Temperature curve (5o resolution, 771  

    from 1900-2000); the transition from ξ(2)/2  ≈  0.4  to  ≈  -‐0.2  occurs  at  τow  ≈  1  yr.    On  the  772  

    lower  left,  we  see  at  daily  resolution,  the  corresponding  globally  averaged  structure  773  

    function.      774  

    Globally  averaged  series:    The  same  20CR  data  but  globally  averaged  (brown),  The  775  

    average  of  the  three  in  situ  surface  series  (NOAA  NCDC,  NASA  GISS,  CRU,  red)  as  well  776  

    as   the   average   of   three   post   2003  multiproxy   structure   functions;   [Huang, 2004],  777  

    [Ljundqvist, 2010; Moberg et al., 2005],  (see  [Lovejoy and Schertzer, 2012c]).      778  

    Paleotemperatures:   At   the   right  we   show   analysis   of   the   EPICA   Antarctic   series  779  

    interpolated   to  50  year   resolution  over  ≈  800  kyrs.    Also   shown   is   the   interglacial  780  

    “window”.  781  

    The reference slopes are ξ(2)/2 = -0.4, -0.2 or +0.4; these correspond to spectral 782  

    exponents β = 1+ξ(2) = 0.2, 0.6, 1.8 respectively; a flat curve in the above corresponds to 783  

    β = 1.  784  

     785  

  • Page 36

     786